首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
An early step in the morphogenesis of the double-stranded DNA (dsDNA) bacteriophage HK97 is the assembly of a precursor shell (prohead I) from 420 copies of a 384-residue subunit (gp5). Although formation of prohead I requires direct participation of gp5 residues 2-103 (Δ-domain), this domain is eliminated by viral protease prior to subsequent shell maturation and DNA packaging. The prohead I Δ-domain is thought to resemble a phage scaffolding protein, by virtue of its highly α-helical secondary structure and a tertiary fold that projects inward from the interior surface of the shell. Here, we employ factor analysis of temperature-dependent Raman spectra to characterize the thermostability of the Δ-domain secondary structure and to quantify the thermodynamic parameters of Δ-domain unfolding. The results are compared for the Δ-domain within the prohead I architecture (in situ) and for a recombinantly expressed 111-residue peptide (in vitro). We find that the α-helicity (∼ 70%), median melting temperature (Tm = 58 °C), enthalpy (ΔHm = 50 ± 5 kcal mol− 1), entropy (ΔSm = 150 ± 10 cal mol− 1 K− 1), and average cooperative melting unit (〈nc〉 ∼ 3.5) of the in situ Δ-domain are altered in vitro, indicating specific interdomain interactions within prohead I. Thus, the in vitro Δ-domain, despite an enhanced helical secondary structure (∼ 90% α-helix), exhibits diminished thermostability (Tm = 40 °C; ΔHm = 27 ± 2 kcal mol− 1; ΔSm = 86 ± 6 cal mol− 1 K− 1) and noncooperative unfolding (〈nc〉 ∼ 1) vis-à-vis the in situ Δ-domain. Temperature-dependent Raman markers of subunit side chains, particularly those of Phe and Trp residues, also confirm different local interactions for the in situ and in vitro Δ-domains. The present results clarify the key role of the gp5 Δ-domain in prohead I architecture by providing direct evidence of domain structure stabilization and interdomain interactions within the assembled shell.  相似文献   

2.
Exchange of DNA between bacteria involves conjugative pili. While the prevailing view has been that F-pili are completely retracted before single-stranded DNA is passed from one cell to another, it has recently been reported that the F-pilus, in addition to establishing the contact between mating cells, serves as a channel for passing DNA between spatially separated cells during conjugation. The structure and function of F-pili are poorly understood. They are built from a single subunit having only 70 residues, and the small size of the subunit has made these filaments difficult to study. Here, we have applied electron cryo-microscopy and single-particle methods to solve the long-existing ambiguity in the packing geometry of F-pilin subunits. We show that the F-pilus has an entirely different symmetry from any of the known bacterial pili as well as any of the filamentous bacteriophages, which have been suggested to be structural homologs. Two subunit packing schemes were identified: one has stacked rings of four subunits axially spaced by ∼ 12.8 Å, while the other has a one-start helical symmetry with an axial rise of ∼ 3.5 Å per subunit and a pitch of ∼ 12.2 Å. Both structures have a central lumen of ∼ 30 Å diameter that is more than large enough to allow for the passage of single-stranded DNA. Remarkably, both schemes appear to coexist within the same filaments, in contrast to filamentous phages that have been described as belonging to one of two possible symmetry classes. For the segments composed of rings, the twist between adjacent rings is quite variable, while the segments having a one-start helix are in multiple states of both twist and extension. This coexistence of two very different symmetries is similar to what has recently been reported for an archaeal Methanococcus maripaludis pili filament and an archaeal Sulfolobus shibatae flagellar filament.  相似文献   

3.
Proteorhodopsin is an ion-translocating member of the microbial rhodopsin family. Light absorption by its retinal chromophore initiates a photocycle, driven by trans/cis isomerization, leading to transmembrane translocation of a proton toward the extracellular side of the cytoplasmic membrane. Here we report a study on the photoisomerization dynamics of the retinal chromophore of proteorhodopsin, using femtosecond time-resolved spectroscopy, by probing in the visible- and in the midinfrared spectral regions. Experiments were performed both at pH 9.5 (a physiologically relevant pH value in which the primary proton acceptor of the protonated Schiff base, Asp97, is deprotonated) and at pH 6.5 (with Asp97 protonated). Simultaneous analysis of the data sets recorded in the two spectral regions and at both pH values reveals a multiexponential excited state decay, with time constants of ∼0.2 ps, ∼2 ps, and ∼20 ps. From the difference spectra associated with these dynamics, we conclude that there are two chromophore-isomerizaton pathways that lead to the K-state: one with an effective rate of ∼(2 ps)−1 and the other with a rate of ∼(20 ps)−1. At high pH, both pathways are equally effective, with an estimated quantum yield for K-formation of ∼0.7. At pH 6.5, the slower pathway is less productive, which results in an isomerization quantum yield of 0.5. We further observe an ultrafast response of residue Asp227, which forms part of the counterion complex, corresponding to a strengthening of its hydrogen bond with the Schiff base on K-state formation; and a feature that develops on the 0.2 ps and 2 ps timescale and probably reflects a response of an amide II band in reaction to the isomerization process.  相似文献   

4.
A current popular model to explain phosphorylation of smooth muscle myosin (SMM) by myosin light-chain kinase (MLCK) proposes that MLCK is bound tightly to actin but weakly to SMM. We found that MLCK and calmodulin (CaM) co-purify with unphosphorylated SMM from chicken gizzard, suggesting that they are tightly bound. Although the MLCK:SMM molar ratio in SMM preparations was well below stoichiometric (1:73 ± 9), the ratio was ∼ 23-37% of that in gizzard tissue. Fifteen to 30% of MLCK was associated with CaM at ∼ 1 nM free [Ca2+]. There were two MLCK pools that bound unphosphorylated SMM with Kd ∼ 10 and 0.2 μM and phosphorylated SMM with Kd ∼ 20 and 0.2 μM. Using an in vitro motility assay to measure actin sliding velocities, we showed that the co-purifying MLCK-CaM was activated by Ca2+ and phosphorylation of SMM occurred at a pCa50 of 6.1 and at a Hill coefficient of 0.9. Similar properties were observed from reconstituted MLCK-CaM-SMM. Using motility assays, co-sedimentation assays, and on-coverslip enzyme-linked immunosorbent assays to quantify proteins on the motility assay coverslip, we provide strong evidence that most of the MLCK is bound directly to SMM through the telokin domain and some may also be bound to both SMM and to co-purifying actin through the N-terminal actin-binding domain. These results suggest that this MLCK may play a role in the initiation of contraction.  相似文献   

5.
Oxygen transport in Myriapoda is maintained by a unique 6 × 6mer hemocyanin, that is, 36 subunits arranged as six hexamers (1 × 6mers). In the sluggish diplopod Spirostreptus, the 1 × 6mers seem to operate as almost or fully independent allosteric units (h ∼ 1.3; P50 ∼ 5 torr), whereas in the swift centipede Scutigera, they intensively cooperate allosterically (h ∼ 10; P50 ∼ 50 torr). Here, we show the chemomechanical basis of this differential behavior as deduced from hybrid 6 × 6mer structures, obtained by single-particle cryo-electron microscopy of the Scutigera 6 × 6mer (10.0 Å resolution according to the 0.5 criterion) and docking of homology-modeled subunits from Scutigera and two diplopods, Spirostreptus and Polydesmus. The Scutigera 6 × 6mer hemocyanin is a trigonal antiprism assembled from six smaller trigonal antiprisms (1 × 6mers), thereby exhibiting D3 point group symmetry. It can be described as two staggered 3 × 6mers or three oblique 2 × 6mers. Topologically, the 6 × 6mer is subdivided into six subunit zones, thereby exhibiting a mantle (24 subunits) and a core (12 subunits). The six hexamers are linked by 21 bridges, subdivided into five types: two within each 3 × 6mer and three between both 3 × 6mers. The molecular models of the 6 × 6mer reveal intriguing amino acid appositions at these inter-hexamer interfaces. Besides opportunities for salt bridges, we found pairs of carboxylate residues for possible bridging via a Ca2+ or Mg2+ ion. Moreover, we detected histidine clusters, notably in Scutigera, allowing us to advance hypotheses as to how the hexamers are allosterically coupled in centipede hemocyanin and why they act more independently in diplopod hemocyanin.  相似文献   

6.
The translocase of the outer mitochondrial membrane (TOM) complex is the main entry gate for proteins imported into mitochondria. We determined the structure of the native, unstained ∼ 550-kDa core-Tom20 complex from Saccharomycescerevisiae by cryo-electron microscopy at 18-Å resolution. The complex is triangular, measuring 145 Å on edge, and has near-3-fold symmetry. Its bulk is made up of three globular ∼ 50-Å domains. Three elliptical pores on the c-face merge into one central ∼ 70-Å cavity with a cage-like assembly on the opposite t-face. Nitrilotriacetic acid-gold labeling indicates that three Tom22 subunits in the TOM complex are located at the perimeter of the complex near the interface of the globular domains. We assign Tom22, which controls complex assembly, to three peripheral protrusions on the c-face, while the Tom20 subunit is tentatively assigned to the central protrusion on this surface. Based on our three-dimensional map, we propose a model of transient interactions and functional dynamics of the TOM assembly.  相似文献   

7.
A subclass of proteins with the SEA (sea urchin sperm protein, enterokinase, and agrin) domain fold exists as heterodimers generated by autoproteolytic cleavage within a characteristic G− 1S+ 1VVV sequence. Autoproteolysis occurs by a nucleophilic attack of the serine hydroxyl on the vicinal glycine carbonyl followed by an N → O acyl shift and hydrolysis of the resulting ester. The reaction has been suggested to be accelerated by the straining of the scissile peptide bond upon protein folding. In an accompanying article, we report the mechanism; in this article, we provide further key evidence and account for the energetics of coupled protein folding and autoproteolysis. Cleavage of the GPR116 domain and that of the MUC1 SEA domain occur with half-life (t½) values of 12 and 18 min, respectively, with lowering of the free energy of the activation barrier by ∼ 10 kcal mol− 1 compared with uncatalyzed hydrolysis. The free energies of unfolding of the GPR116 and MUC1 SEA domains were measured to ∼ 11 and ∼ 15 kcal mol− 1, respectively, but ∼ 7 kcal mol− 1 of conformational energy is partitioned as strain over the scissile peptide bond in the precursor to catalyze autoproteolysis by substrate destabilization. A straining energy of ∼ 7 kcal mol− 1 was measured by using both a pre-equilibrium model to analyze stability and cleavage kinetics data obtained with the GPR116 SEA domain destabilized by core mutations or urea addition, as well as the difference in thermodynamic stabilities of the MUC1 SEA precursor mutant S1098A (with a G− 1A+ 1VVV motif) and the wild-type protein. The results imply that cleavage by N → O acyl shift alone would proceed with a t½ of ∼ 2.3 years, which is too slow to be biochemically effective. A subsequent review of structural data on other self-cleaving proteins suggests that conformational strain of the scissile peptide bond may be a common mechanism of autoproteolysis.  相似文献   

8.
Favourable mutations involving the two dehalogenases (DehI and DehII) of Pseudomonas putida PP3 and derivative strains containing the cloned gene for DehI (dehI) occurred in response to specific environmental conditions, namely: starvation conditions; the presence of dehalogenase substrates (halogenated alkanoic acids — HAAs) which were toxic to P. putida; and/or the presence of a potential growth substrate. Fluctuation tests showed that these mutations were environmentally directed by the presence of HAAs. the mutations were associated with complex DNA rearrangements involving the movement of dehI located on a transposon DEH. Some mutations resulted in switching off the expression of either one or both of the dehalogenases, events which were effective in protecting P. putida from toxic compounds in its growth environment. Other mutations partially restored P. putida's dehalogenating capability under conditions where toxic substrates were absent. Restoration of the capability to untilize HAAs was favoured when normal growth substrates were present in the environment.  相似文献   

9.
Several RNA-cleaving deoxyribozymes (DNAzymes) have been reported for efficient cleavage of purine-containing junctions, but none is able to efficiently cleave pyrimidine-pyrimidine (Pyr-Pyr) junctions. We hypothesize that a stronger Pyr-Pyr cleavage activity requires larger DNAzymes with complex structures that are difficult to isolate directly from a DNA library; one possible way to obtain such DNAzymes is to optimize DNA sequences with weak activities. To test this, we carried out an in vitro selection study to derive DNAzymes capable of cleaving an rC-T junction in a chimeric DNA/RNA substrate from DNA libraries constructed through chemical mutagenesis of five previous DNAzymes with a kobs of ∼ 0.001 min− 1 for the rC-T junction. After several rounds of selective amplification, DNAzyme descendants with a kobs of ∼ 0.1 min− 1 were obtained from a DNAzyme pool. The most efficient motif, denoted “CT10-3.29,” was found to have a catalytic core of ∼ 50 nt, larger than other known RNA-cleaving DNAzymes, and its secondary structure contains five short duplexes confined by a four-way junction. Several variants of CT10-3.29 exhibit a kobs of 0.3-1.4 min− 1 against the rC-T junction. CT10-3.29 also shows strong activity (kobs  > 0.1 min− 1) for rU-A and rU-T junctions, medium activity (> 0.01 min− 1) for rC-A and rA-T junctions, and weak activity (> 0.001 min− 1) for rA-A, rG-T, and rG-A junctions. Interestingly, a single-point mutation within the catalytic core of CT10-3.29 altered the pattern of junction specificity with a significantly decreased ability to cleave rC-T and rC-A junctions and a substantially increased ability to cleave rA-A, rA-T, rG-A, rG-T, rU-A, and rU-T junctions. This observation illustrates the intricacy and plasticity of this RNA-cleaving DNAzyme in dinucleotide junction selectivity. The current study shows that it is feasible to derive efficient DNAzymes for a difficult chemical task and reveals that DNAzymes require more complex structural solutions for such a task.  相似文献   

10.
Effects of grinding processes on enzymatic degradation of wheat straw   总被引:1,自引:0,他引:1  
The effectiveness of wheat straw fine to ultra-fine grindings at pilot scale was studied. The produced powders were characterised by their particle-size distribution (laser diffraction), crystallinity (WAXS) and enzymatic degradability (Trichoderma reesei enzymatic cocktail). A large range of wheat-straw powders was produced: from coarse (median particle size ∼800 μm) to fine particles (∼50 μm) using sieve-based grindings, then ultra-fine particles ∼20 μm by jet milling and ∼10 μm by ball milling. The wheat straw degradability was enhanced by the decrease of particle size until a limit: ∼100 μm, up to 36% total carbohydrate and 40% glucose hydrolysis yields. Ball milling samples overcame this limit up to 46% total carbohydrate and 72% glucose yields as a consequence of cellulose crystallinity reduction (from 22% to 13%). Ball milling appeared to be an effective pretreatment with similar glucose yield and superior carbohydrate yield compared to steam explosion pretreatment.  相似文献   

11.
A good understanding of cell membrane properties is crucial for better controlled and reproducible experiments, particularly for cell electroporation where the mechanism of pore formation is not fully elucidated. In this article we study the influence on that process of several constituents found in natural membranes using bilayer lipid membranes. This is achieved by measuring the electroporation threshold (Vth) defined as the potential at which pores appear in the membrane. We start from highly stable 1,2-diphytanoyl-sn-glycero-3-phosphocholine (DPhPC) membranes (Vth ∼ 200 mV), and subsequently add therein other phospholipids, cholesterol and a channel protein. While the phospholipid composition has a slight effect (100 mV ≤ Vth ≤ 290 mV), cholesterol gives a concentration-dependent effect: a slight stabilization until 5% weight (Vth ∼ 250 mV) followed by a noticeable destabilization (Vth ∼ 100 mV at 20%). Interestingly, the presence of a model protein, α-hemolysin, dramatically disfavours membrane poration and Vth shows a 4-fold increase (∼ 800 mV) from a protein density in the membrane of 24 × 10− 3 proteins/μm2. In general, we find that pore formation is affected by the molecular organization (packing and ordering) in the membrane and by its thickness. We correlate the resulting changes in molecular interactions to theories on pore formation.  相似文献   

12.
The tightly packaged double-stranded DNA (dsDNA) genome in the mature particles of many tailed bacteriophages has been shown to form multiple concentric rings when reconstructed from cryo-electron micrographs. However, recent single-particle DNA packaging force measurements have suggested that incompletely packaged DNA (ipDNA) is less ordered when it is shorter than ∼ 25% of the full genome length. The study presented here initially achieves both the isolation and the ipDNA length-based fractionation of ipDNA-containing T3 phage capsids (ipDNA-capsids) produced by DNA packaging in vivo; some ipDNA has quantized lengths, as judged by high-resolution gel electrophoresis of expelled DNA. This is the first isolation of such particles among the tailed dsDNA bacteriophages. The ipDNA-capsids are a minor component (containing ∼ 10− 4 of packaged DNA in all particles) and are initially detected by nondenaturing gel electrophoresis after partial purification by buoyant density centrifugation. The primary contaminants are aggregates of phage particles and empty capsids. This study then investigates ipDNA conformations by the first cryo-electron microscopy of ipDNA-capsids produced in vivo. The 3-D structures of DNA-free capsids, ipDNA-capsids with various lengths of ipDNA, and mature bacteriophage are reconstructed, which reveals the typical T = 7l icosahedral shell of many tailed dsDNA bacteriophages. Though the icosahedral shell structures of these capsids are indistinguishable at the current resolution for the protein shell (∼ 15 Å), the conformations of the DNA inside the shell are drastically different. T3 ipDNA-capsids with 10.6 kb or shorter dsDNA (< 28% of total genome) have an ipDNA conformation indistinguishable from random. However, T3 ipDNA-capsids with 22 kb DNA (58% of total genome) form a single DNA ring next to the inner surface of the capsid shell. In contrast, dsDNA fully packaged (38.2 kb) in mature T3 phage particles forms multiple concentric rings such as those seen in other tailed dsDNA bacteriophages. The distance between the icosahedral shell and the outermost DNA ring decreases in the mature, fully packaged phage structure. These results suggest that, in the early stage of DNA packaging, the dsDNA genome is randomly distributed inside the capsid, not preferentially packaged against the inner surface of the capsid shell, and that the multiple concentric dsDNA rings seen later are the results of pressure-driven close-packing.  相似文献   

13.
The type III secretion system (T3SS) is essential for the infectivity of many pathogenic Gram-negative bacteria. The T3SS contains proteins that form a channel in the inner and outer bacterial membranes, as well as an extracellular needle that is used for transporting and injecting effector proteins into a host cell. The homology between the T3SS and the bacterial flagellar system has been firmly established, based upon both sequence similarities between respective proteins in the two systems and the structural homology of higher-order assemblies. It has previously been shown that the Shigella flexneri needle has a helical symmetry of ∼ 5.6 subunits/turn, which is quite similar to that of the most intensively studied flagellar filament (from Salmonella typhimurium), which has ∼ 5.5 subunits/turn. We now show that the Sa. typhimurium needle, expected by homology arguments to be more similar to the Sa. typhimurium flagellar filament than is the needle from Shigella, actually has ∼ 6.3 subunits/turn. It is not currently understood how host cell contact, made at the tip of the needle, is communicated to the secretory system at the base. In contrast to the Sa. typhimurium flagellar filament, which shows a nearly crystalline order, the Sa. typhimurium needle has a highly variable symmetry, which could be used to transmit information about host cell contact.  相似文献   

14.
Each difference between the genome sequences of Escherichia coli B strains REL606 and BL21(DE3) can be interpreted in light of known laboratory manipulations plus a gene conversion between ribosomal RNA operons. Two treatments with 1-methyl-3-nitro-1-nitrosoguanidine in the REL606 lineage produced at least 93 single-base-pair mutations (∼ 90% GC-to-AT transitions) and 3 single-base-pair GC deletions. Two UV treatments in the BL21(DE3) lineage produced only 4 single-base-pair mutations but 16 large deletions. P1 transductions from K-12 into the two B lineages produced 317 single-base-pair differences and 9 insertions or deletions, reflecting differences between B DNA in BL21(DE3) and integrated restriction fragments of K-12 DNA inherited by REL606. Two sites showed selective enrichment of spontaneous mutations. No unselected spontaneous single-base-pair mutations were evident. The genome sequences revealed that a progenitor of REL606 had been misidentified, explaining initially perplexing differences. Limited sequencing of other B strains defined characteristic properties of B and allowed assembly of the inferred genome of the ancestral B of Delbrück and Luria. Comparison of the B and K-12 genomes shows that more than half of the 3793 proteins of their basic genomes are predicted to be identical, although ∼ 310 appear to be functional in either B or K-12 but not in both. The ancestral basic genome appears to have had ∼ 4039 coding sequences occupying ∼ 4.0 Mbp. Repeated horizontal transfer from diverged Escherichia coli genomes and homologous recombination may explain the observed variable distribution of single-base-pair differences. Fifteen sites are occupied by phage-related elements, but only six by comparable elements at the same site. More than 50 sites are occupied by IS elements in both B and K, 16 in common, and likely founding IS elements are identified. A signature of widespread cryptic phage P4-type mobile elements was identified. Complex deletions (dense clusters of small deletions and substitutions) apparently removed nonessential genes from ∼ 30 sites in the basic genomes.  相似文献   

15.
Using cryo-electron microscopy, single particle image processing and three-dimensional reconstruction with icosahedral averaging, we have determined the three-dimensional solution structure of bacteriophage MS2 capsids reassembled from recombinant protein in the presence of short oligonucleotides. We have also significantly extended the resolution of the previously reported structure of the wild-type MS2 virion. The structures of recombinant MS2 capsids reveal clear density for bound RNA beneath the coat protein binding sites on the inner surface of the T = 3 MS2 capsid, and show that a short extension of the minimal assembly initiation sequence that promotes an increase in the efficiency of assembly, interacts with the protein capsid forming a network of bound RNA. The structure of the wild-type MS2 virion at ∼9 Å resolution reveals icosahedrally ordered density encompassing ∼90% of the single-stranded RNA genome. The genome in the wild-type virion is arranged as two concentric shells of density, connected along the 5-fold symmetry axes of the particle. This novel RNA fold provides new constraints for models of viral assembly.  相似文献   

16.
The bindings of biogenic polyamines spermine (spm), spermidine (spmd) and synthetic polyamines 3,7,11,15-tetrazaheptadecane·4HCl (BE-333) and 3,7,11,15,19-pentazahenicosane·5HCl (BE-3333) with β-lactoglobulin (β-LG) were determined in aqueous solution. FTIR, UV-vis, CD and fluorescence spectroscopic methods as well as molecular modeling were used to determine the polyamine binding sites and the effect of polyamine complexation on protein stability and secondary structure. Structural analysis showed that polyamines bind β-LG via both hydrophilic and hydrophobic contacts. Stronger polyamine-protein complexes formed with synthetic polyamines than biogenic polyamines, with overall binding constants of Kspm-β-LG = 3.2(±0.6) × 104 M−1, Kspmd-β-LG = 1.8(±0.5) × 104 M−1, KBE-333-β-LG = 5.8(±0.3) × 104 M−1 and KBE-3333-β-LG = 6.2(±0.05) × 104 M−1. Molecular modeling showed the participation of several amino acids in the polyamine complexes with the following order of polyamine-protein binding affinity: BE-3333 > BE-333 > spermine > spermidine, which correlates with their positively charged amino group content. Alteration of protein conformation was observed with a reduction of β-sheet from 57% (free protein) to 55-51%, and a major increase of turn structure from 13% (free protein) to ∼21% in the polyamine-β-LG complexes, indicating a partial protein unfolding.  相似文献   

17.
18.
19.
At room temperature, the chlorophyll (Chl) a fluorescence induction (FI) kinetics of plants, algae and cyanobacteria go through two maxima, P at ∼ 0.2-1 and M at ∼ 100-500 s, with a minimum S at ∼ 2-10 s in between. Thus, the whole FI kinetic pattern comprises a fast OPS transient (with O denoting origin) and a slower SMT transient (with T denoting terminal state). Here, we examined the phenomenology and the etiology of the SMT transient of the phycobilisome (PBS)-containing cyanobacterium Synechococcus sp PCC 7942 by modifying PBS → Photosystem (PS) II excitation transfer indirectly, either by blocking or by maximizing the PBS → PS I excitation transfer. Blocking the PBS → PS I excitation transfer route with N-ethyl-maleimide [NEM; A. N. Glazer, Y. Gindt, C. F. Chan, and K.Sauer, Photosynth. Research 40 (1994) 167-173] increases both the PBS excitation share of PS II and Chl a fluorescence. Maximizing it, on the other hand, by suspending cyanobactrial cells in hyper-osmotic media [G. C. Papageorgiou, A. Alygizaki-Zorba, Biochim. Biophys. Acta 1335 (1997) 1-4] diminishes both the PBS excitation share of PS II and Chl a fluorescence. Here, we show for the first time that, in either case, the slow SMT transient of FI disappears and is replaced by continuous P → T fluorescence decay, reminiscent of the typical P → T fluorescence decay of higher plants and algae. A similar P → T decay was also displayed by DCMU-treated Synechococcus cells at 2 °C. To interpret this phenomenology, we assume that after dark adaptation cyanobacteria exist in a low fluorescence state (state 2) and transit to a high fluorescence state (state 1) when, upon light acclimation, PS I is forced to run faster than PS II. In these organisms, a state 2 → 1 fluorescence increase plus electron transport-dependent dequenching processes dominate the SM rise and maximal fluorescence output is at M which lies above the P maximum of the fast FI transient. In contrast, dark-adapted plants and algae exist in state 1 and upon illumination they display an extended P → T decay that sometimes is interrupted by a shallow SMT transient, with M below P. This decay is dominated by a state 1 → 2 fluorescence lowering, as well as by electron transport-dependent quenching processes. When the regulation of the PBS → PS I electronic excitation transfer is eliminated (as for example in hyper-osmotic suspensions, after NEM treatment and at low temperature), the FI pattern of Synechococcus becomes plant-like.  相似文献   

20.
The GroEL/GroES protein folding chamber is formed and dissociated by ATP binding and hydrolysis. ATP hydrolysis in the GroES-bound (cis) ring gates entry of ATP into the opposite unoccupied trans ring, which allosterically ejects cis ligands. While earlier studies suggested that hydrolysis of cis ATP is the rate-limiting step of the cycle (t½ ∼ 10 s), a recent study suggested that ADP release from the cis ring may be rate-limiting (t½ ∼ 15-20 s). Here we have measured ADP release using a coupled enzyme assay and observed a t½ for release of ?4-5 s, indicating that this is not the rate-limiting step of the reaction cycle.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号